Sie sind auf Seite 1von 9

Applied Physics Research; Vol. 8, No.

2; 2016
ISSN 1916-9639 E-ISSN 1916-9647
Published by Canadian Center of Science and Education

Computation of the Coefficients of the Power law model for Whole


Blood and Their Correlation with Blood Parameters
Mohamed A. Elblbesy1,2 & Abdelrahman T. Hereba1
1
Department of Medical Biophysics, Medical Research Institute, Alexandria University, Egypt
2
Department of Medical Laboratory Technology, Faculty of Applied Medical Science, University of Tabuk, Saudi
Arabia
Correspondence: Mohamed A. Elblbesy, Department of Medical Biophysics, Medical Research Institute,
Alexandria University, Egypt. Email:mimizizo@yahoo.com, melblbesy@ut.edu.sa

Received: January 4, 2016 Accepted: January 15, 2016 Online Published: February 15, 2016
doi:10.5539/apr.v8n2p1 URL: http://dx.doi.org/10.5539/apr.v8n2p1

Abstract
This study introduces a quantitative analysis of the coefficients of the power law model, which is used to
describe the non-Newtonian behavior of blood. Twenty blood samples from healthy donors were used to
measure the whole blood viscosity under different values of the shear rates, which are between 2.25 and 450.0 s-1.
The shear rate viscosity curves were used to calculate n (flow index) and m (the consistency of the fluid)
according to the power law model. Strong correlations (R2 > 0.5) between m and the hematocrit (HCT %),
hemoglobin (Hb), erythrocytes count (RBC), mean corpuscle volume (MCV), and mean corpuscle hemoglobin
concentration (MCHC) were obtained. Strong correlations (R2 > 0.5) between n and the RBC, MCV, and MCHC
were achieved. The relation obtained between the power law coefficients and the blood parameters in the present
investigation provides new parameters that can be used to evaluate the flow state of blood besides blood
viscosity. In addition, these parameters may be used to examine blood under pathological conditions,
representing a new tool for the diagnosis of blood abnormal conditions.
Keywords: Blood, Viscosity, Power law, HCT, Hb, RBC, MCV, MCHC
1. Introduction
Blood is a heterogeneous multi-phase mixture of solid corpuscles (red blood cells, white blood cells, and
platelets) suspended in liquid plasma, which is an aqueous solution of proteins, organic molecules, and minerals.
The rheological characteristics of blood are determined by the properties of these components and their
interaction with each other as well as with the surrounding structures. Blood rheology is also affected by external
physical conditions, such as temperature; however, in living organisms, in general, and in large mammals, in
particular, these conditions are regulated, and hence, they are subject to minor variations that cannot significantly
affect the general properties (Bodnar, 2011). Other physical properties, such as the mass density, may also play a
role in determining the overall blood rheological conduct. The rheological properties of blood and blood vessels
are affected by the intake of fluids, nutrients, and medications, although in most cases, these effects are not
substantial, except possibly over short periods of time, and normally do not have lasting consequence
(Breithaupt-Grogler et al., 1997; Vlastos et al., 2003; Marcinkowska Gapinska et al., 2007; Tripette et al., 2010).
Plasma behaves as a Newtonian fluid, whereas whole blood has non-Newtonian properties. In large vessels,
where shear the rates are high, it is reasonable to assume that blood has a constant viscosity and a Newtonian
behavior. However, in smaller vessels, or in some disease conditions, the presence of cell induces a low shear
rate and whole blood exhibits remarkable non-Newtonian characteristics, such as shear-thinning viscosity,
thixotropy, viscoelasticity and possibly a yield stress. This is largely due to the behavior of RBCs, namely, their
ability to aggregate into microstructures (rouleau) at low shear rates, their deformability into an infinite variety
of shapes without changing volume and their tendency to align with the flow field at high shear rates (Chien et
al., 1970; Schmid-Schenbein & Wells, 1969). An understanding of the coupling between the blood composition
and the physical properties of blood is essential to develop suitable constitutive models to describe blood
behavior (see the recent reviews) (Robertson et al., 2009; Robertson et al., 2008).

1
www.ccsenet.org/apr Applied Physics Research Vol. 8, No. 2; 2016

Many models have been developed to describe blood viscosity. One of the most widely used forms of the general
non-Newtonian constitutive relation is a power-law model, which can be described as (Bird et al., 1987):
= (1)
Where is shear stress, is shear rate and m, n are power-law model constants. The constant m is a measure
of the consistency of the fluid: the higher the value of m, the more viscous the fluid is. n is a measure of the
degree of non-Newtonian behavior: the greater the departure from the unity, the more pronounced the
non-Newtonian properties of the fluid are. Viscosity for the power-law fluid can be expressed as (Bird R B. et al.,
1987):
= (2)
where is the non-Newtonian apparent viscosity. If n < 1, then a shear thinning fluid is obtained; such a fluid is
characterized by a progressively decreasing apparent viscosity with an increasing shear rate (Bird et al., 1987).
The present work aims to compute n and m for whole blood and test their relationship with the physiological
properties of blood. In addition, this work explores whether these power-law model constants reflect the state of
the internal structure and the interaction between blood contents.
2. Materials and Method
2.1 Sample Collection
Twenty blood samples were collected from donors of the same age and gender. All blood samples were collected
after overnight fasting in a quiet environment at normal ambient temperature. Blood was withdrawn after a 10
minute resting period and in a seated position. Blood was collected from the antecubital vein 90 s after application
of a tourniquet and without removing it to avoid the effect of pressure on the blood samples (Baskurt et al., 2009).
EDTA (1.5 mg/ml) was used as an anticoagulant. Two milliliters from each sample was sent to the clinical
laboratory to measure the hematocrit (HCT %), hemoglobin (Hb), erythrocytes count (RBC), mean corpuscle
volume (MCV), and mean corpuscle hemoglobin concentration (MCHC) using a hematology auto analyzer.
2.2 Blood Viscosity Measurement
Viscosity measurements were performed at 37 oC using a rotational viscometer fitted with a small sample cup
and a temperature controller (Brookfield Digital LVTD viscometer). The viscometer was calibrated according to
the manufacturers explicit, step-by-step instructions. Proper calibration and operation were verified through the
measurement of the viscosity of the standard solution for both Newtonian and non-Newtonian fluids.
The viscosities were measured at the different shear rates specified on a Brookfield viscometer: 2.25, 4.50, 11.25,
22.5, 45.0, 90.0, 225.0 and 450.0 sec-1. A volume of 0.5 ml of a whole blood sample was added to the viscometer.
For each test, the measurement was repeated ten times, and then, the average of the readings was taken. Each test
was performed on a new sample. The cone, cup, and spindle assembly were rinsed and carefully dried between
each trial.
2.3 Power Law Models Coefficient Calculations
The relationship between log and log was a straight line according to the power law model. log was the
y-axis and log was the x-axis. n and m were calculated graphically. n was calculated from the slope of the line
( = 1). m was calculated from the intercept of the line with the y-axis (log =
). Microsoft Excel 2010 was used to calculate both n and m. Linear regression was used after the
relationship between log and log was plotted. The equation of the linear trend line was used to compute n
and m, as mentioned above.
2.4 Statistical Analysis
The results are presented as the mean SD unless otherwise noted. Correlations were made using Pearsons
coefficient. Pearson's R2 can range from -1 to 1. A R2of -1 indicates a perfect negative linear relationship between
variables, a R2 of 0 indicates no linear relationship between variables, and a R2 of 1 indicates a perfect positive
linear relationship between variables. The data within each group were analyzed using Microsoft Excel 2010.
3. Results
The viscosity data revealed the non-Newtonian behavior of normal human blood over the given range of shear
rates (Figure 1). Shear thinning of whole blood was viewed through the relationship between and (Figure
1). A sharp decrease in was observed between 2.25 and 450.0 sec-1. The measured values of were 12.5
mPa.s at a low shear rate (2.25 sec-1), 4.89 mPa.s at 45 sec-1, and 2.38 mPa.s at a high shear rate (450.0 sec-1).

2
www.ccsenet.org/apr Applied Physics Research Vol. 8, No. 2; 2016

The logarithmic relationship between and was represented as a straight line that intercepted the y-axis
(Figure 2). This relation was used to compute the values of m and n. n was in the range of 0.71630.00145 to
0.66720.00178, and m was in the range of 16.240.0064 mPa.sn to 16.1010.003 mPa.sn
n was found to be strongly correlated with RBC and MCV (R2>0.5). A positive correlation was indicated for the
relationship between n and RBC. By contrast, negative correlations were indicated for the relation between n and
HTC and MCV (Figure 3). A moderate positive correlation was found between n and MCHC (R2=0.5). A positive
weak correlation was found between n and Hb (R2<0.5). By contrast, a negative weak correlation was found
between n and HTC (Figure 3).
m was strongly correlated to RBC, Hb, HCT, MCV, and MCHC (R2>0.5). The relationships between n and the
RBC, Hb, and MCHC were inversely proportional, whereas n and HCT and MCV were directly proportional.

Figure 1. Whole blood viscosity as a function of the shear rate shows non-Newtonian behavior. Whole blood
shear thinning over the range of shear rates from 2.25 to 450 sec-1

Figure 2. The logarithmic relationship between viscosity and the shear rate shows a high degree of linearity. The
intercept with the y-axis is used to calculate m, and the slope of the line is used to calculate n

3
www.ccsenet.org/apr Applied Physics Research Vol. 8, No. 2; 2016

(A) (B)
0.73 0.73

0.72 0.72

0.71 0.71

0.7 0.7
n

n
0.69 0.69

0.68 0.68

0.67 0.67

0.66 0.66
4 4.2 4.4 4.6 4.8 12 13 14 15 16 17
RBC Hb (g/dL)
(C) (D)
0.73
0.73

0.72 0.72

0.71 0.71

0.7 0.7
n
n

0.69 0.69

0.68 0.68

0.67 0.67

0.66 0.66
39 40 41 42 43 44 80 85 90 95 100 105

HTC % MCV (fL)


(E)

0.73

0.72

0.71

0.7
n

0.69

0.68

0.67

0.66
30 32 34 36 38 40 42

MCHC (g/dL)

Figure 3. n correlates to some blood parameters. (A) The equation for the relationship between n and the RBC
count is n = 0.0834 RBC + 0.3218 and R = 0.61. (B) The equation for the relationship between n and
Hb is n = 0.0104 Hb + 0.5468 and R = 0.42. (C) The equation for the relationship between n and HTC
is n = 0.0114 HTC + 1.1617 and R = 0.32. (D) The equation for the relationship between n and MCV
is n = 0.0026 MCV + 0.9362 and R = 0.6. (E) The equation for the relationship between n and MCHC is
n = 0.0039 MCHC + 0.5579 and R = 0.5

4
www.ccsenet.org/apr Applied Physics Research Vol. 8, No. 2; 2016

(A) (B)
16.26 16.26

16.24 16.24

16.22 16.22

16.2 16.2
m (mPa.sn)

m (mPa.sn)
16.18 16.18
16.16 16.16
16.14 16.14
16.12 16.12
16.1 16.1
16.08 16.08
4 4.2 4.4 4.6 4.8 12 13 14 15 16 17
RBC * 106 Hb (g/dL)
(C) (D)

16.26 16.26
16.24 16.24
16.22 16.22
16.2 16.2
m (mPa.sn)
m (mPa.sn)

16.18 16.18

16.16 16.16

16.14 16.14

16.12 16.12

16.1 16.1

16.08 16.08
39 40 41 42 43 44 80 85 90 95 100 105
HCT % MCV (fL)

16.26
16.24 (E)
16.22
16.2
m (mPa.sn)

16.18
16.16
16.14
16.12
16.1
16.08
30 35 40 45
MCHC (g/dL)

Figure 4. Correlation of m to some blood parameters. (A) The equation for the relationship between m and the
RBC count is m = 0.2006 RBC + 17.058 and R = 0.63. (B) The equation for the relationship
between m and Hb is m = 0.0302 Hb + 16.592 and R = 0.64. (C) The equation for the relationship
between m and HTC is m = 0.0359 HTC + 14.689 and R = 0.6. (D) The equation for the relationship
between m and MCV is m = 0.0069 MCV + 15.522 and R = 0.71. (E) The equation for the relationship
between m and MCHC is m = 0.0116 MCHC + 16.565 and R = 0.75

5
www.ccsenet.org/apr Applied Physics Research Vol. 8, No. 2; 2016

4. Discussion
Blood has previously been indicated to be a non-Newtonian fluid and exhibits a complex rheological behavior,
such as shear-thinning viscosity and thixotropy, primarily due to the presence of and interaction between cellular
elements, mainly red blood cells (RBCs), which are the most abundant component and whose mechanical
properties are inherent to the microstructure characteristic of blood (Long et al., 2005; Gijsen et al., 1999;
Thurston, 1972; Sousa et al., 2013). The non-Newtonian behavior of blood is mainly explained by three
phenomena: the tendency of erythrocytes to form three-dimensional microstructures (rouleau) at low shear rates,
their deformability (or breakup), and their tendency to align with the flow field at high shear rates. Chien et al.
indicated that when blood is at rest or at low shear rates (below 1 s-1), it seems to have a high apparent viscosity,
whereas at high shear rates, there is a reduction in the blood's viscosity (Chien et al., 1967). The blood viscosity
decreased with an increasing shear rate. For sufficiently high shear rates, the viscosity reached a plateau. Data
(minimum of 132 data points) were fitted to a curve that decayed as a monotonic exponential in response to an
increasing shear rate (David et al., 2000). The mean blood viscosity obtained by Robert S et al. at shear rates of
100, 50, and 1 s-1 were 3.26 0.43,4.37 0.60, and 5.46 0.84 mPa.s, respectively (Robert et al., 1996). The
blood viscosity was measured using a Coulter-Harkners viscometer at high shear rates (over 300 s -1). The
normal controls correspond to blood viscosity values of 3.53 mPa.s and a corrected blood viscosity of 3.46 mPa.s
(Lowe et al., 1993). Other studies reported that at low shear rates or shear stresses, the apparent viscosity was
high, whereas the apparent viscosity decreased with increasing shear and approached a minimum value under
high shear forces. At high shear rates above 100 to 200 sec-1, the viscosity of normal blood measured at 37C is
approximately 4 to 5 mPa.s and is relatively insensitive to further increases of shear. However, the viscosity
becomes increasingly sensitive to shear rates below 100 s-1 and increases exponentially as the shear rate is
decreased. Nominal values for the viscosity of normal blood are approximately 10 mPa.s at 10 s-1, 20 mPa.s at 1
s-1, and 100 mPa.s at 0.1 s-1. Thus, at lower shear rates, the blood viscosity becomes extremely sensitive to the
decrement in shear forces. At stasis, normal blood has a yield stress in the range of 2 to 4 mPa (Merrill EW, 1969;
Rampling, 1988; Chien., 1975; Rand et al., 1964). Our results (Figure 1.), obtained using a rotational viscometer,
are in good accord with the results obtained in the literature for both the non-Newtonian behavior of whole blood
and for the values of the apparent blood viscosity in the given range of shear rates. The difference in the apparent
viscosity at a specific shear rate, especially under low shear rates, from previous studies is due to the difference
in the hematocrit percentages used in our study.
Experiments with the shear-rheometer in a Couette geometry (Behbahani et al., 2009) were used to
accommodate the Parabolic model to obtain experimental data for the shear rate and shear stress. Six rheological
measurements of human blood with 47 % hematocrit were analyzed, and the results for the Parabolic model are
compared with the widely used Power law model. For the latter parameters, K (consistency index) and n (Flow
index) were determined using the least-squares method (Marn & Ternik, 2003). The values of K were in the
range of 1.568E-2 Pa.s to 1.8567E-2 Pa.s, and the values of n were between 7.4815E-1 and 7.18E-1. Bernasconi
et al. and Kar et al. have shown that the power law regression method is suitable for blood viscosity
quantification. Bernasconi et al. conducted their series of experiments on exclusively normal blood samples; Kar
et al. and later Hussain et al., from the same research group, conclusively proved that not only normal blood but
also other pathological blood types follow a power law model (Mohammad et al., 1990). The results obtained are
extremely useful for understanding the rheological behavior of blood. A few studies determined the correlation
of the power-law coefficients with the hematological and biochemical parameters in the form of the constitutive
equation. Easthope and Brooks used a constitutive function that was first employed by Walburn and Schneck to
describe the flow properties of whole blood that relate the shear stress measured in a viscometer to the shear rate
and hematocrit of the sample (Walburn & Schneck, 1976; Easthope & Brooks, 1976). Hussain et al. reported that
because n of the power law model is the non-Newtonian behavior index and k is the flow consistency index of
blood, they are naturally dependent on the constituents of blood, such as hematocrit, fibrinogen, cholesterol, and
so on. It is possible to have such a relationship between the power law coefficients and the above-mentioned
parameters in the form of a mathematical equation by using non-linear regression analysis (Mohammad et al.,
1990; Hussain et al., 1994; Hussain et al., 1995). Panagiotis used m = 14.67 mPa.s and n = 0.7755 for his
simulation of blood flow; this simulation represented a typical behavior of blood flow according to the power
law model (Panagiotis Neofytou, 2004). For healthy controls, Hussain et al. reported that the experimental value
of n was 0.708 and the calculated value of n was 0.713, with 0.25 as the standard deviation of the difference
between the experimental and calculated values. Alternatively, the value of normalized K for healthy controls
(divided by 17.0 mPa.sn) was 0.980 and 0.995 for the experimental and calculated values, respectively, with
0.106 as the standard deviation of the difference between the experimental and calculated values (Mohammad et
al., 1990). The calculated n and K from the research of Hussain et al. were calculated as functions of hematocrit,

6
www.ccsenet.org/apr Applied Physics Research Vol. 8, No. 2; 2016

fibrinogen, and cholesterol; their results proved that n and K are correlated strongly with the selected blood
parameters (Mohammad et al., 1990). Blood viscosity can be affected by various factors (e.g., hemospherine,
glucose, and proteins), the most important of which is hematocrit, which denotes the percentage of blood volume
that is occupied by red blood cells (Beatriz et al., 2006). Whole blood viscosity correlated with HCT (r=0.63, p <
0.001 at low shear and r =0.84, p < 0.001 at high shear) (Craig et al., 2006). Karsheva et al. proved the
dependence of the power law coefficients on both RBC and HCT (M. Karsheva et al., 2009). They obtained the
value of n<1 and m between 0.9 to 1.3 mPa.sn. A key feature of our study is to examine the correlations between
m and n with the blood parameters that are mainly related to erythrocytes (such as HCT %, Hb, RBC, MCV, and
MCHC) because erythrocytes play a major role in the non-Newtonian behavior of blood. The values of n and m
obtained from our study are in a good agreement with previous studies. The correlations between n and both
RBC and MCV were strong (Figure 3), in accordance with the dramatic dependence of the number and size of
erythrocytes on the flow behavior of the whole blood. The correlations between n and other blood indices were
moderate to weak (Figure 3) because the other blood parameters have stronger effects on the physiological and
biochemical properties of whole blood than the macro-rheological behavior of whole blood.
RBC and their related indices are well known to affect viscosity (Chao-Hung, 2004). Chao-Hung found strong
positive correlations between whole blood viscosity and the RBC count, Hb, and HCT. David M. et al. found
that the blood viscosity increased with increasing hematocrit (David et al., 2000). M, which depends on the
physiological parameters of blood, was strongly correlated to the blood indices considered in this study (Figure
4). HCT and Hg increase blood viscosity under different values of shears rates. This linear logarithmic
relationship between viscosity and the shear rates corresponds to lines with higher slopes and increasing
intercepts of the line, with the y-axis representing the shear rates. This leads to an increase of the value of blood
indices with increasing m. From the dependence of m on the internal structure of the liquid under investigation,
our results showed strong positive correlations with all of the blood indices.
5. Conclusion
From this study, we conclude that blood behaves as a non-Newtonian fluid, exhibiting a shear thinning behavior.
Many models could be used to describe the non-Newtonian behavior of blood, the easiest of which is the power
law model. Much information could be extracted from the power law model using an easy computational method.
This information is related to the physiological parameters of blood; as a result, the extraction of information
from the power law model is a powerful tool to evaluate blood under normal and abnormal conditions. The
power law coefficients, n and m, used in this study and in previous studies showed promise as a method to
describe the state of blood under investigation based on the correlations between the coefficients and blood
parameters (such as HCT, Hb, RBC, MCV in this study and fibrinogen and cholesterol in other study). Further
research should be performed to study the correlation between the power law coefficients and the blood
physiological and chemical parameters under both normal and pathogenic conditions. In addition, research
should be performed to evaluate the values of the power law model coefficients for blood at low shear rates.
References
Baskurt, O., Boynard, M., Cokelet, G., Connes, P., Cooke, B. M., Forconi, S., ... & Nash, G. (2009). New
guidelines for hemorheological laboratory techniques. Clinical hemorheology and microcirculation, 42(2),
75-97.
Behbahani, M., Behr, M., Hormes, M., Steinseifer, U., Arora, D., Coronado, O., & Pasquali, M. (2009).
Computational Fluid Dynamics of Blood Flow. European Journal of applied mathematics, 20, 363-397.
http://dx.doi.org/10.1017/S0956792509007839
Bird, R. B., Curtiss, C. F., Armstrong, R. C., & Hassager, O. (1987). Dynamics of polymeric liquids. In Kinetic
theory (Volume 2, p. 437). New York: Wiley-lnterscience.
Bodnr, T., Sequeira, A., & Prosi, M. (2011). On the shear-thinning and viscoelastic effects of blood flow under
various flow rates. Applied Mathematics and Computation, 217(11), 5055-5067. http://dx.doi.org/10.
1016/j.amc.2010.07.054
Breithaupt-Grgler, K., Ling, M., Boudoulas, H., & Belz, G. G. (1997). Protective effect of chronic garlic intake
on elastic properties of aorta in the elderly. Circulation, 96(8), 2649-2655. http://dx.doi.org/10.
1161/01.CIR.96.8.2649
Broberg, C. S., Bax, B. E., Okonko, D. O., Rampling, M. W., Bayne, S., Harries, C., ... & Gibbs, J. S. R. (2006).
Blood viscosity and its relationship to iron deficiency, symptoms, and exercise capacity in adults with
cyanotic congenital heart disease. Journal of the American College of Cardiology, 48(2), 356-365.

7
www.ccsenet.org/apr Applied Physics Research Vol. 8, No. 2; 2016

Chien, S. (1975). Biophysical behavior of red cells in suspensions. The red blood cell, 2(4), 1031-1133.
http://dx.doi.org/10.1016/B978-0-12-677202-9.50019-8
Chien, S., Usami, S., Dellenback, R. J., & Gregersen, M. I. (1970). Shear-dependent deformation of erythrocytes
in rheology of human blood. American Journal of Physiology--Legacy Content, 219(1), 136-142.
Chien, S., Usami, S., Dellenback, R. J., Gregersen, M. I., Nanninga, L. B., & Guest, M. M. (1967). Blood
viscosity: influence of erythrocyte aggregation. Science, 157(3790), 829-831. http://dx.doi.org/10.1126/
science.157.3790.829
Easthope, P. L., & Brooks, D. E. (1979). A comparison of rheological constitutive functions for whole human
blood. Biorheology, 17(3), 235-247.
Eckmann, D. M., Bowers, S., Stecker, M., & Cheung, A. T. (2000). Hematocrit, volume expander, temperature,
and shear rate effects on blood viscosity. Anesthesia & Analgesia, 91(3), 539-545. http://dx.doi.org/10.
1097/00000539-200009000-00007
Gijsen, F. J. H., Allanic, E., Van de Vosse, F. N., & Janssen, J. D. (1999). The influence of the non-Newtonian
properties of blood on the flow in large arteries: unsteady flow in a 90 curved tube. Journal of biomechanics,
32(6), 601-608. http://dx.doi.org/10.1016/S0021-9290(99)00015-9
Ho, C. H. (2004). White blood cell and platelet counts could affect whole blood viscosity. Journal-Chinese
Medical Association, 67(8), 394-397.
Hussain, M. A., Kar, S., & Puniyani, R. R. (1999). Relationship between power law coefficients and major blood
constituents affecting the whole blood viscosity. Journal of Biosciences, 24(3), 329-337.
Hussain, M. A., Puniyani, R. R., & Kar, S. (1994). Quantification of blood viscosity using power law model in
cerebrovascular accidents and high risk controls. Clin. Hemorheo L., 14, 685-696.
Hussain, M. A., Puniyani, R. R., & Kar, S. (1995). Blood viscosity autoregulation factor in cerebrovascular
accidents and high risk cases-> a new approach. Clinical hemorheology, 1(15), 61-71.
Karsheva, M., Dinkova, P., Pentchev, I., & Ivanova, T. (2009). Blood rheology-a key for blood circulation in
human body. Journal of the University of Chemical Technology and Metallurgy, 44(1), 50-54.
Long, J. A., ndar, A., Manning, K. B., & Deutsch, S. (2005). Viscoelasticity of pediatric blood and its
implications for the testing of a pulsatile pediatric blood pump. ASAIO journal, 51(5), 563-566.
http://dx.doi.org/10.1097/01.mat.0000180353.12963.f2
Lowe, G. D., Fowkes, F. G., Dawes, J., Donnan, P. T., Lennie, S. E., & Housley, E. (1993). Blood viscosity,
fibrinogen, and activation of coagulation and leukocytes in peripheral arterial disease and the normal
population in the Edinburgh Artery Study. Circulation, 87(6), 1915-1920. http://dx.doi.org/10.1161/
01.CIR.87.6.1915
Marcinkowska-Gapiska, A., Gapinski, J., Elikowski, W., Jaroszyk, F., & Kubisz, L. (2007). Comparison of
three rheological models of shear flow behavior studied on blood samples from post-infarction patients.
Medical & biological engineering & computing, 45(9), 837-844. http://dx.doi.org/10.1007/s11517
-007-0236-4
Marn, J., & Ternik, P. (2003). Use of Quadratic model for modelling of fly ash-water mixture. Applied Rheology,
13(6), 286-296.
Merrill, E. W. (1969). Rheology of blood. Physiol. Rev, 49(4), 863-888.
Neofytou, P. (2003). Comparison of blood rheological models for physiological flow simulation. Biorheology,
41(6), 693-714.
Rampling, M. W. (1988). Red cell aggregation and yield stress. Clinical blood rheology, 1, 11-44.
Rand, P. W., Lacombe, E., Hunt, H. E., & Austin, W. H. (1964). Viscosity of normal human blood under
normothermic and hypothermic conditions. Journal of Applied Physiology, 19(1), 117-122.
Robertson, A. M., Sequeira, A., & Kameneva, M. V. (2008). Hemorheology. In Hemodynamical Flows
Modeling, Analysis and Simulation, Birkhuser Basel (pp. 63-120).
Robertson, A. M., Sequeira, A., & Owens, R. G. (2009). Rheological models for blood. In Cardiovascular
mathematics (pp. 211-241). Springer Milan. http://dx.doi.org/10.1007/978-88-470-1152-6_6

8
www.ccsenet.org/apr Applied Physics Research Vol. 8, No. 2; 2016

Rosenson, R. S., McCormick, A., & Uretz, E. F. (1996). Distribution of blood viscosity values and biochemical
correlates in healthy adults. Clinical Chemistry, 42(8), 1189-1195. http://dx.doi.org/10.1016/0021
-9150(94)94059-2
Salazar-Vazquez, B. Y., Intaglietta, M., Rodrguez-Morn, M., & Guerrero-Romero, F. (2006). Blood pressure
and hematocrit in diabetes and the role of endothelial responses in the variability of blood viscosity.
Diabetes Care, 29(7), 1523-1528. http://dx.doi.org/10.2337/dc06-0323
Schmid-Schnbein, H., & Wells, R. (1969). Fluid drop-like transition of erythrocytes under shear. Science,
165(3890), 288-291.
Sousa, P. C., Carneiro, J., Vaz, R., Cerejo, A., Pinho, F. T., Alves, M. A., & Oliveira, M. (2014). Shear viscosity
and nonlinear behaviour of whole blood under large amplitude oscillatory shear. Biorheology, 50(5-6),
269-282.
Thurston, G. B. (1972). Viscoelasticity of human blood. Biophysical journal, 12(9), 1205-1217.
http://dx.doi.org/10.1016/S0006-3495(72)86156-3
Tripette, J., Loko, G., Samb, A., Gogh, B. D., Sewade, E., Seck, D., ... & Brudey, K. (2010). Effects of hydration
and dehydration on blood rheology in sickle cell trait carriers during exercise. American Journal of
Physiology-Heart and Circulatory Physiology, 299(3), H908-H914. http://dx.doi.org/10.1152/ajpheart.
00298.2010
Vlastosa, G. A., Tangneyb, C. C., & Rosensonc, R. S. (2003). Effects of hydration on blood rheology. Clinical
hemorheology and microcirculation, 28, 41-49.
Walburn, F. J., & Schneck, D. J. (1976). A constitutive equation for whole human blood. Biorheology, 13(3),
201-210.

Copyrights
Copyright for this article is retained by the author(s), with first publication rights granted to the journal.
This is an open-access article distributed under the terms and conditions of the Creative Commons Attribution
license (http://creativecommons.org/licenses/by/3.0/).

Das könnte Ihnen auch gefallen